You are on page 1of 35

7

Better Unification for Physics in General


Through Quantum Mechanics in Particular
Cynthia Kolb Whitney
Galilean Electrodynamics
USA

1. Introduction
Physics has always had several different domains of application in on-going development,
and physicists have always striven for unification among its different domains. Unification
is usually achieved through development of so-called ‘covering theories’. In the nineteenth
century, the stunning example was Maxwell’s Electrodynamics (MED), which unified
electricity and magnetism as one domain of theory. Another major domain of theory then
present was Newton’s Mechanics (NM), which in the eighteenth century had really
launched modern physics as a mathematical discipline.
At the turn of the twentieth century, NM and MED were well in place, and were fulfilling
many technologically important requirements. But there seemed to be an incompatibility
between them. The problem concerned their invariance with respect to choice of reference
frame: NM exhibited invariance if the allowed reference frames were all connected through
Galilean transformations, whereas MED exhibited invariance if the allowed reference frames
were all connected through Lorentz transformations. It looked as though one of these two
theories must be more nearly correct than the other, but it was not clear which one was the
better one.
That problem seemed resolved with the advent of Einstein’s Special Relativity Theory (SRT).
SRT was believed to capture the true meaning of MED concerning the behavior of light
signals, and SRT was certainly an endorsement of Lorentz transformation, so SRT was
believed to offer the one possible revision of NM that could make mechanics fully consistent
with MED.
But meanwhile, new phenomena were being discovered at the micro scale of physics, and
they often seemed inexplicable with any known theory, whether NM, SRT, or MED. These
were phenomena suggesting quantization of light, quantized atomic states, atomic,
molecular and crystal structures, radioactivity, etc.
So at almost the same time as one problem seemed to be resolved, other problems were
emerging. Since the earlier situation between NM and MED had demanded that Physics allow
two seemingly discordant theories to co-exist until some good argument could replace one of
them, the situation then presented by the new phenomena being discovered naturally invited
the development of another potentially discordant theory: Quantum Mechanics (QM).
The discovery of the photoelectric effect, and the introduction of the idea of the photon,
initiated QM. Almost immediately, QM was developed to handle the Hydrogen atom, and
the ground state thereof, the stability of which was thought to be impossible with MED.

www.intechopen.com
128 Theoretical Concepts of Quantum Mechanics

Accepting that apparent incompatibility with MED, and even embracing it, researchers
moved on to excited states, to other atoms, then to molecules, and reactions, and to all the
rest of the complexity that today makes up modern Quantum Chemistry (QC).
Also, experimenters got into sub-atomic elementary particles, especially electrons and
positrons, their annihilation and creation, along with creation and annihilation of photons.
All that led to Quantum Electrodynamics (QED).
So today physics still has several different bodies of theory, aimed at several different
domains of application. On the one hand, we have QM for atomic and other micro-system
interactions. It has at least two identifiable parts: QC for interactions at the level of atoms
and molecules, and QED for interactions at the level of elementary particles. And on the
other hand, we have Einstein’s relativity theory (RT) for physics at human scale and larger.
It too has two parts: SRT for electromagnetic interactions, and general relativity theory
(GRT) for gravitational interactions.
QM and RT are the major pillars of twentieth century physics. And they are not entirely
compatible. QM features wave-like entities with seemingly instantaneous correlations
between the states of even quite distant entities, whereas RT features point-like entities
interacting via fields propagating at a finite speed.
So are we defeated in the quest for unification in Physics? Apparently many people hope
not, as they do vigorously pursue various forms of unification. The prominent one sought
today is Quantum Gravity (QG). It would be the twenty-first century capstone for the two
twentieth-century pillars of QM and RT. But it is not yet fully in sight.
In the pursuit of unification, one often sees phrases like ‘Theory of Everything’. The
objective of this Chapter is certainly modest by comparison! It just notes some observations
about the status of available theories, and discusses the removal of some incompatibilities
between the available theories that arose only because of unfortunate choices.
Because QM is relatively new, there are still lots of alternative approaches being developed
in parallel. Putz (2009) gives us one very big and recent anthology about them, and this
book will give another even more recent one. The QM atmosphere is clearly right for
generating new illumination that can facilitate new observations about physics overall.
The first observation driving the present work is just this: QED is arguably the most
successful theory that modern Physics possesses. The fact that QED now exists, and that is
has the name that it has, naturally begs the question: How could there have been any real
disconnect between MED1 and early QM?
It is this author’s belief that Nature is not so perverse. Connections between different
domains of theory are still possible to find, even though the diligent search that was
conducted a century ago did not find them. We have developed more tools now. Every new
tool developed should invite us to revisit the old problems.
Section 2 talks about the photon from the point of view of MED. It explores the implications
of the finite energy, which characterizes a photon. It finds a plausible model for the photon
expressed in terms of MED.
The second observation is just this: If MED can connect better with QM, then shouldn’t SRT
also connect better with QM? After all, how much difference can there be between a photon
in QM and a light signal in SRT?

1 Note that I speak of MED, not of Classical Electrodynamics (CED) in general. CED involves, not only

the works of Maxwell, but also those of a large number of other individuals. I am inclined to trust
results from Maxwell, but question some of those from other authors, as reported in the present work.

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 129

Section 3 explores the implications of modeling the light signal in SRT in the same way as
the photon in QM. The photon model suggests a slight alteration to Einstein’s second
postulate, and thereby produces a slightly altered version of SRT.
The third observation is just this: If SRT is to be altered, however slightly, in response to the
photon concept from QM, isn’t it then possible that the revised SRT can be used to better
explain some things about QM that presently seem mysterious?
Section 4 talks about what the photon/signal model implies about atoms: the stability of
atoms, the occurrence of Planck’s constant.
The fourth observation is this: Much of science works on scaling laws. It is in that spirit that
we should look for scaling laws about atoms, and thereby reduce the effort of looking at
each element as a particular special-case problem for detailed calculations.
Section 5 talks about the inferences from to the story about all isotopes of Hydrogen, all
elements beyond Hydrogen, and the ions of any element; the possible nature of ‘excited’
atomic states, and the character of the light spectrum that an element produces.
The fifth observation is this: If QM can be better connected to SRT, then where does that
leave its relationship with NM? Early QM was basically NM, although not for particles
possessing momentum and energy in the classical way, but rather for waves, with an
amplitude factor and a phase factor, in the latter of which momentum and energy appeared
as variables. Is that formulation now completely outdated on account of a rift between NM
and MED?
Section 6 establishes that there was no necessary disconnect even between NM and MED. It
argues that, with an adequately extended notation to support an extended tensor calculus,
Maxwell’s equations can be seen to be invariant in form, even under Galilean transformation.
(It is useful here to distinguish two kinds of invariance: ‘form invariance’ for symbolic
equations, and ‘number invariance’ for individual symbols that have numerical values.)
The last observation is the ‘meta’ observation about the present work: Physics in general can
become significantly more unified throughout because of some specific developments
surrounding QM.
Section 7 summarizes the several specific conclusions implied by the present work. Boiled
down to one sentence, these conclusions come to this: the existence of apparent discord
between theories that are addressed to different problem domains within Physics sometimes
means that there exists a more productive way to pose one or more of the theories involved.

2. Maxwell’s electrodynamics and QM’s photons


It often seems that MED, a theory largely about spatially extended EM fields, has little in
common with QM, a theory largely about discrete material systems and the discrete photons
that they emit and absorb. Photons are imagined to be the opposite of spatially extended;
i.e., localized, like the matter particles that emit and absorb them.
So our mental picture for a photon in its interactions with matter is rather bullet-like: the
photon is shot out of a source, travels through space, and hits a receiver that absorbs it. But
the travel part of the story is unobservable. So we imagine that the photon in flight is
possibly wavelike, in accord with Maxwell theory. Certainly the evidence for that is present,
in the form of interference effects, even with small numbers of photons. So the photon is
assigned a quality of ‘duality’. This is a rather mysterious way of describing a photon.
What seems missing here is an adequate model for the photon throughout its life history,
expressed in terms of EM fields. The purpose of this Section is to develop one.

www.intechopen.com
130 Theoretical Concepts of Quantum Mechanics

I like to begin the development of such a history with a waveform consisting of finite energy
distributed in a three-dimensional Gaussian peak located very close to a source that has
emitted it. This three-dimensional Gaussian peak is limited in all three spatial directions so
as to integrate to a finite total energy.
To allow subsequent propagation, the energy has to be divided between two orthogonal
fields, electric and magnetic. To allow circular polarization, the energy has to be further
divided between real and imaginary parts, real being alive now, and imaginary becoming
alive a quarter of an oscillation cycle later.
Given such a start, the whole life history of a photon can then develop in the manner that
Maxwell’s equations allow. Describing that development is the objective of the following
Sub-Sections.

2.1 Waveform development


The first step in the life history of a photon is its development from a spatially localized
energy bundle that is emitted from a source into a spatially extended waveform that travels
through space. To help think about this problem, it is useful to recall some phenomenology
familiar from physics at a more macroscopic scale.
1. One phenomenon very well known for light modeled as EM waves is the spreading
transverse to the propagation direction known as of ‘diffraction’. Diffraction is the
result of some sort of limitation transverse to the propagation direction. Historically, the
limitation has been due to a finite aperture through which the light propagates. The
light spreads out from the aperture, more-so the smaller the aperture is. In the photon
model discussed here, the limitation is softer than an aperture edge, but a limitation
nevertheless: it is the finite spread of the Gaussian waveform in the two directions
transverse to the propagation direction. The more narrow the Gaussian peak is, the
more spread there will be.
But sideways spreading is not the main requirement for a photon model; spreading in
the longitudinal direction is what is most needed. Could longitudinal spreading be
caused in a manner similar to diffraction, by the initial waveform limitation in the
longitudinal direction?
2. The closest familiar analog for longitudinal spreading is known as ‘dispersion’. This
word refers to the ‘blurring’ effect that any frequency dependence the propagation
speed through the medium entails. For example, a signal pulse in a medium looses its
sharp edges because those sharp edges imply superposition of many different
wavelengths, and hence different frequencies, which the medium may affect differently.
In Earth’s atmosphere, or ocean, square waves can turn to blob waves because of
dispersion.
But we don’t have the traditional medium-induced frequency dispersion for a photon
in free space. So ‘dispersion’ isn’t a close analog for any effect that may be induced by
longitudinal limitation due to the finite spread of the Gaussian waveform in the
longitudinal direction.
For the photon model, we need to find and combine just the useful features from both
the diffraction and dispersion ideas. Here is a workable approach. Diffraction comes
out of optical system response in the spatial domain. Dispersion comes out of
transmission system response in the temporal domain. Maxwell’s equations link space
and time variation together. So we look at pulse profiles in the longitudinal direction,
and allow Maxwell’s equations to work on them.

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 131

propagation direction, say x , with E  exp(  x 2 ) . We can apply Maxwell’s equations, and
Let us begin a scenario with a single pulse in E . Let it have a Gaussian profile along the

watch what happens. The Gaussian is the so-called ‘generating function’ for the infinite set
of Hermite polynomials, all of which have very regularly spaced zero crossings. What
happens is that the single pulse in E (an even function) generates a double pulse in B (an
odd function), which in turn generates a triple pulse in E (another even function), and so

polynomials multiplying the Gaussian. Meanwhile, all the E  B Poynting vectors in play
on; that is, all the derivatives in play generate successively higher-order Hermite

support general spreading of the Gaussian. With each step, the emergent functions look
more and more like wavelets, and the individual peaks in the wavelets stay about the same
width as more of them accrue, so the wavelength for the emergent wavelet becomes more
and more defined. Figure 1 illustrates this behavior at the stage where E has developed five
peaks (four zero crossings). Series 1 is the original input Gaussian function, Series 2 is the
Gaussian after the overall spreading has developed to this point, and Series 3 is the wavelet
that has emerged in the process; i.e. the spread-out Gaussian times the fourth-order Hermit
polynomial generated.

0.8

0.6

0.4

Series1
0.2 Series2
Series3

0
1 4 7 10 13 16 19 22 25 28 31 34 37 40 43 46 49 52 55 58 61 64 67 70 73 76 79

-0.2

-0.4

-0.6

Fig. 1. A wavelet develops when an EM pulse is acted upon by Maxwell’s equations.


What we have so far is only one eighth of the story needed to fully represent a photon:
development from a pulse into a waveform. We have told the story for one pulse in E . If we
would match that with another pulse in B , we would have overall propagation along with

photon. If we would match that with two more pulses, E and B pointing at 90 in space
waveform development. That would bring us to one quarter of the whole story of the

from the first pair and coming ‘alive’ a quarter cycle out of phase with the first pair, we
would have the circular polarization characteristic of photons, but we would still have just
half the story. So let us move on, and seek the other half.

www.intechopen.com
132 Theoretical Concepts of Quantum Mechanics

2.2 Waveform regression


The remaining half of the story of the photon is about waveform regression. How does this
complex structure of four Hermite polynomials multiplied by their generating Gaussian
unwind, and go back to being a set of four pulses, so that it can be absorbed into a receiver?
Again, let us refer to some similar but more familiar phenomenology:
3. A third phenomenon possible for light modeled as EM waves is ‘focusing’. This is what
we have optical lenses and shaped mirrors for. It works somewhat contrary to
transverse spreading, gathering incident energy into a smaller area transverse to the
propagation direction. Of course we don’t have any lenses or mirrors in the photon
model, but we shall find a mechanism that produces a similar effect.
4. A fourth phenomenon possible for light modeled as EM waves is ‘pulse restoration’.
This is what transmission lines have ‘repeater stations’ for. A communication signal
degraded by dispersion can be reconstituted when passed through an intelligent filter.
Of course we don’t have any filters in a photon model, but we shall find a mechanism
that produces a similar effect.
The ‘similar effect’ comes from the imposition of boundary conditions in the longitudinal
direction. The Gaussian pulse that was used to describe the waveform development part of
the scenario was somewhat unrealistic in that its tails extended to infinity. There is no way
that a localized source could emit an energy pulse whose tails would extend to infinity. It is
somewhat more realistic to imagine the equivalent of a mirror at the source, and another
mirror at the eventual receiver, to confine the waveform like a wave in a box, with zero
amplitude at the surface of each mirror and everywhere beyond.
With such boundary conditions imposed, the analytic functions involved in the model are
no longer the simple Gaussian and the simple Hermite polynomials that it generates. Now
we have not one, but three, Gaussians, the extra two being needed to cancel the first one at
the two boundaries. Correspondingly, we always have at least three (actually six) Hermite
polynomials alive at any given time. That is a loss of mathematical simplicity. But there is a
gain of conceptual simplicity. It is easy to envision that the propagation scenario has some
symmetry about its mid point. The waveform will spread until its central peak is halfway
between the source and the receiver. After that, the mirror at the receiver will be more
significant than the mirror at the source, causing the waveform to start ‘piling up’ near the
receiver, and eventually end up as a pulse near the receiver, similar to the pulse originally
launched near the source.
This ‘regressing waveform’ is somewhat reminiscent of ‘advanced’ solutions to Maxwell’s
equations going backwards in time. These were introduced many times in the early 20th
century, but particularly popularized in the mid 20th century by Wheeler and Feynman
(1945 and 1949).2 What we have here is quite different though. There are no differential

2 Wheeler and Feynman were looking to time symmetry as the basis for an electromagnetic
generalization of instantaneous (Newtonian) gravitational interaction. There are important differences
between the regressing waveforms introduced above and the Wheeler-Feynman advanced solutions: 1)
Wheeler and Feynman were looking at interactions between essentially point sources and receivers, and
so had to be looking at spherically expanding retarded solutions and spherically contracting advanced
solutions, not at essentially one-dimensional expanding and contracting wavelets. 2) The Wheeler-
Feynman expansion or contraction is related to the spherical area of a wave front, not the waveform in
the radial propagation direction. 3) A lengthy discussion of the paradox of advanced actions is
necessitated in the Wheeler-Feynman work, whereas the ‘regressing’ solutions introduced here are not
in fact ‘advanced’ at all; they are just regressing, in real time, in the propagation direction.

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 133

equations running backwards in time; there is just ‘piling up’ of a solution to differential
equations in response to a boundary condition.

2.3 The photon model in terms of EM fields


Taken together, the waveform development followed by the waveform regression suggest a
photon model in terms of EM fields that exhibits continuous evolution: it goes from a state
of pulse-like localization near its source, to a state of wave-like extension in space during its
travel, and then back to a state of pulse-like localization near its receiver.
Observe that with this photon model, ‘light in flight’ develops its wavelength only during
its flight. It doesn’t have it to start with, and it gives it up at the end. So light at emission, or
reception, has a position, but no wavelength, whereas light in flight has a wavelength, but
no position. Thus the model expresses a ‘wave-particle duality’ for light.
Observe too that this photon model exhibits a form of QM ‘complementarity’, or uncertainty
relationship. Consider that, under Fourier transformation, Gaussians map into Gaussians,
and that the product of the spreads of such Gaussians is a constant. In the process of wave
train development, a Gaussian in position space x spreads out, while its corresponding
Gaussian in wave number space k sharpens up.
Inasmuch as the discovery of photons was the point of departure for the development of
QM, having this photon model expressed in terms of Maxwell fields is a first step in
reconciling MED with QM. But there is much more to do, because the bigger problem for
MED was not the photon itself, but rather the atom that emitted or absorbed it. It looked as
though MED could never explain an atom being stable in its ground state, much less
anything about its excited states. To find any reconciliation there, we must move on.

3. EM signals as photons
Every neutral atom contains at least two particles, and generally a lot more. Prior to QM,
electromagnetic forces were presumed to hold such a system together, but there was clearly
a problem with that understanding.
The simplest atom is the Hydrogen atom, with just one electron circulating about a nucleus
consisting of just one proton. So consider the Hydrogen atom. The electron circulates and so
accelerates, and that must generate radiation. It was assumed that this radiation would rob
the atomic system of energy, and thereby cause the collapse of the atom.
So it was assumed that Maxwell’s EMT is simply incompatible with the stability of atoms.
The solution then was to postulate the existence of a different regime of physics in which
that wouldn’t happen. But was that really necessary? The purpose of this Section is to argue
that it was not.
The underlying belief in inevitability of atomic collapse reflects a belief that the
electrodynamic forces within the atom are essentially central, and therefore cannot affect the
energy budget of the atom. This latter belief traces to the turn of the 20th century, when A.
Liénard (1898) and E. Wiechert (1901) developed models for the potentials and fields created
by rapidly moving charges. Although Liénard and Wiechert worked independently, they
made the same assumption, and they got the same results, and so confirmed each other.
This Section looks at those results, and thereby develops a motivation to look back at their
underlying assumption.

www.intechopen.com
134 Theoretical Concepts of Quantum Mechanics

3.1 Standard formulae for scalar and vector potentials


Expressed in Gaussian units, the Liénard-Wiechert (LW) scalar and vector potentials at
position r and time t are

( r , t )  e  1 /  R retarded and A( r , t )  e   /  R 

(1)
retarded

where   1  n   ,  is source velocity normalized by c , and n  R / R (a unit vector),


 

and R  rsource (t  R / c )  r (an implicit definition for the terminology ‘retarded’). The LW
fields obtained from those potentials are then

 ( n   )(1   2 )   d  
 
E( x , t )  e         and B( r , t )  n retarded  E( r , t )
n
  3R2 c 3 R  dt  
( n ) (2)
retarded

The LW fields have some interesting properties. The 1 / R fields are radiation fields, and
they make a Poynting vector (energy flow per unit area per unit time) that lies along
n retarded :

P Eradiative  Bradiative  Eradiative   n retarded  Eradiative  


4 4 4
c c c
(Eradiative )2 n retarded (3)

But the 1 / R 2 fields are Coulomb-Ampère fields, and the Coulomb


n retarded as one might naively expect; instead, it lies along ( n   )retarded . Assume
 that 
 field does not lie along

does not change much over the total field propagation time, in which case ( n   )retarded is
virtually indistinguishable from n present . So then the Coulomb field and the radiation are
arriving to the observer from different directions.
One can feel moved to check this surprising result. Fortunately, one can look up the original
sources, obtain translations if necessary, and verify the original algebra. There is no problem
with the algebra. There are also numerous re-derivations that use more modern techniques
involving the Dirac delta function and the Heaviside step function. These are ‘generalized’
functions of some parameter that, when driven to infinity, produces an infinite pulse or a
unit step. One can study these re-derivations too. One finds various re-orderings of the
mathematical operators ‘differentiate’, ‘integrate’, and ‘go to parameter limit’. These re-
orderings are dodgy because the generalized functions lack the mathematical property of
uniform convergence, so these operations don’t necessarily commute; it is possible to
change the result by changing operation order. But even so, such findings do not change the
fact that the original LW derivations, although pedestrian, were correct.
If a problem exists with this LW result, then there is really only one place where it can arise:
in the initial assumption; namely, that electromagnetic fields propagate like bullets shot at
speed c . But this is the very same assumption that Einstein later formalized as his Second
Postulate (1905, 1907). He just called them “signals” rather than “fields”.
The LW idea of bullets shot at speed c is the foundation for Special Relativity Theory (SRT).
(Indeed, SRT offers one of the modern ways to re-derive the LW results.) But SRT is also the
foundation for General Relativity Theory (GRT). SRT and GRT together make one of the two
great pillars of 20th century Physics: Relativity Theory (RT). So questioning the LW

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 135

assumption is not just questioning the LW results; it is questioning the founding assumption
of SRT, and so threatening this whole pillar of 20th century theory.
Many people have just accepted that this is just ‘the way things are’ with classical field
theory, and with SRT, and with all of relativity theory as well. But what if one wanted to
describe the same scenarios in a thoroughly modern way, with photons instead of radiation
fields, and virtual photons instead of Coulomb-Ampère fields? Could anyone really accept
the idea that the real photons and the virtual photons created by the same space-time event
would arrive at a detector from different directions?
But one needn’t accept any such thing, given the photon model in terms of Maxwell fields
developed in Sect. 2. In short, since we have a model for photons in terms of fields, we
should be able to reverse engineer a model for fields in terms of photons. So what does the
photon model developed in Sect. 2 imply? Observe that the developing wavelet can move at
speed c relative to the source, and the regressing wavelet can move at speed c relative to
the receiver. Applying this idea can help to modify the LW results appropriately.

3.2 Updated formulae for scalar and vector potentials


Recall that with the photon model developed in terms of Maxwell fields in Sect. 2, the life
history of the photon has a symmetry point in the middle. Before the mid point of the
propagation scenario, the waveform is developing, and after the mid point of the
propagation scenario the waveform is regressing. That makes the mid point very important.
So far as the receiver is concerned, nothing that happened before the midpoint affects the
signal he receives. The source position and velocity information he receives is determined,
not by the specification ‘retarded’, but rather by the specification ‘half retarded’.
With this new specification, the scalar and vector potentials become:

( r , t )  e  1 /  Rhalf retarded and A(r, t )  e   /  R 



(4)
half retarded

The fields become:

 ( n   )(1   2 )   d  
 
E( r , t )  e        
n
  3R2 c 3 R  dt  
( n )
half retarded
(5)

and B( r , t )  n half retarded  E( r , t )

The Poynting vector P( r , t ) becomes:

Eradiative  Bradiative  Eradiative   n half retarded  Eradiative 


4 4
c c
(6)

4
c
(Eradiative )2 n half retarded

Observe that now the direction of the Coulomb field is


( n   )half retarded  ( n present )half retarded  n half retarded and the direction of the Poynting vector

is n half retarded too. So now, the Coulomb field and the Poynting vector are reconciled to the

www.intechopen.com
136 Theoretical Concepts of Quantum Mechanics

same direction. That is the first big gift from the photon model in terms of EM fields given in
Sect. 2.
And the gifts of photon model in terms of EM fields go well beyond this rather arcane
problem about field direction. The photon model in terms of EM fields eliminates the central
mystery of Einstein’s SRT: having just one light speed relative to however many different
observers there may be. This is complexity at the level of ‘multiplicity’, much more daunting
than the complexity at the level of the mere ‘duality’ that is found in modern QM.

4. EM fields within atoms


An noted in Sect. 2, atoms were the really big problem for Maxwell’s EMT. Now armed with
some new information about EMT, it is appropriate the revisit the problem about atoms. We
turn again to Hydrogen. From Sect. 3 can infer that at least two processes go on inside the
Hydrogen atom, and we shall discover shortly that there are actually three. Only one is
familiar. The other two challenge familiar concepts of ‘conservation’ that originally grew out
of Newtonian mechanics. But electromagnetism is not Newtonian mechanics. In
electromagnetic problems, the concepts of momentum and energy ‘conservation’ have to
include the momentum and energy of fields, as well as those of matter. Momentum and
energy can both be exchanged between matter and fields. ‘Conservation’ applies only to the
system overall, not to matter alone (nor to fields alone either).

4.1 Energy loss due to far-field radiation


The first process that occurs with the Hydrogen atom is the familiar energy loss from the
atom due to far-field radiation. There will be a far-field power radiated (energy loss per unit
time) of magnitude

  d 
Pradiated   |P|R 2 d   ( Eradiative )2 R 2 d    
 2
      d ,
c e2
4 4 c 2 6
c
 dt 
n ( n ) (7)
4 4 4

where  means ‘solid angle’. Because the full 4 of solid angle captures opposing
directions of n , contributions to the integral from the vector  visible in
cancel out. Contributions to the integral that come from the dot product n   that is hidden
 the integrand

in the  6 factor may not be zero at every moment, but they time-average
simplify the expression for far-field power radiated by setting  to zero. We have:
 to zero. So let us

 d 
4 c 4
 2
Pradiated      d ,
e2
 dt 
n n (8)

It evaluates to the well-known Larmor result:

d e 2 d
4 c 1   d cos (1  cos
2 2 
 d cos d sin   )
1 2 1
e2 2 2
Pradiated
1

 cos  
0
dt 2c dt

e 2 d 2 e 2 d
  . (9)
 cos 3  
2 2
1
1
2c dt 3 1 3c dt

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 137

4.2 Energy gain due to internal torquing


The second process that occurs in the Hydrogen atom is a not previously noticed energy

atom is not central; it is along n half retarded , and not along ( n   )retarded  n present .
gain due to internal torquing. This process occurs because the  Coulomb force within the

The power inflow to the electron is Ptorquing  Te e , where e is the electron orbit
frequency, and Te is the magnitude of the torque on the electron, given by Te  re  Fe
where re is the electron orbit radius, and Fe is the tangential force on the electron. But that
is not all. The proton also orbits at frequency e , and experiences its own torque, given by
Tp  rp  Fp , where rp is the proton orbit radius (tiny) and Fp is the tangential force on the
proton (huge), with the result that the magnitude Tp is the same as Te . The total torque on
the system is T  Te  Tp  2Te . It is determined by the angle between re and Fe , which is
given by rpe / 2c  ( me / mp )re e / 2c . So torque T  (me / mp )(re e / c )e 2 (re  rp ) and
power received is

me re e2
Ptorqing   ( e 4 / mp ) c( re  rp )3 .
e2
mp c (re  rp )
(10)

The existence of such a process is why the concept of ‘balance’ emerges: there can be a
balance between gain of energy due to internal torquing and the inevitable loss of energy
due to radiation. But we are not done with radiation yet.

4.3 Extra radiation due to Thomas rotation


The fact that the electron and the proton have such different masses, and orbit at such
different radii, means that the EM forces within the atom are not only not central; they are
not even balanced. This situation has another major implications: The system as a whole
experiences a net force. That means the system center of mass (C of M) can move. This sort
of effect does not occur in Newtonian mechanics due to the fact that Newtonian mechanics
assumes infinite signal propagation speed.
Looking in more detail, the unbalanced forces in the Hydrogen atom must cause the C of M
of the whole atom to traverse its own circular orbit, on top of the orbits of the electron and
proton individually. This is an additional source of accelerations, and hence of radiation. It
evidently makes even worse the original problem of putative energy loss by radiation that
prompted the development of QM. But on the other hand, the torque on the system is a
candidate mechanism to compensate the rate of energy loss due to radiation, even if there is
a lot more radiation than originally thought..
The details are worked out quantitatively as follows. First ask what the circulation can do to
the radiation. Some 20 years after the advent of SRT, a relevant kinematic truth about
systems traversing circular paths was uncovered by L.H. Thomas (1927), in connection with
explaining the then anomalous magnetic moment of the electron: 1/2 its expected value. He
showed that a coordinate frame attached to a particle driven around a circle naturally
rotates at half the imposed circular revolution rate.
Applied to the scenario of the electron orbiting the proton, the gradually rotating x , y
coordinate frame of the electron means that the electron sees the proton moving only half as
fast as an external observer would see it. That fact explained the electron’s anomalous
magnetic moment, and so was received with great interest in its day. But the fact of Thomas
rotation has since slipped to the status of mere curiosity, because Dirac theory has replaced

www.intechopen.com
138 Theoretical Concepts of Quantum Mechanics

it as the favored explanation for the magnetic moment problem. Now, however, there is a
new problem in which to consider Thomas rotation: the case of the C of M of a whole
Hydrogen atom being driven in a circle by unbalanced forces. In this scenario, the gradually
rotating local x , y coordinate frame of the C of M means that the atom system doing its
internal orbiting at frequency e relative to the C of M will be judged by an external
observer to be orbiting twice as fast, at frequency   2 e relative to inertial space. This

electron-magnetic-moment problem; 2) by construction from e in the C of M system as the


perhaps surprising result can be established in at least three ways: 1) by analogy to the old

power series   e  (1  2  4  8  ...)  e  2 ; 3) by observation that in inertial space


1 1 1

 must satisfy the algebraic relation   e  2  , which implies   2 e .


1

The relation   2 e means the far field radiation power, if it really ever manifested itself

formula for radiation power from a charge e ( e in electrostatic units) is P  2 e 2 a 2 / 3c 3 ,


in the far field, would be even stronger than classically predicted. The classical Larmor

would be from the electron, orbiting with ae  re e2 , e given by the Coulomb force
where a is total acceleration. For the classical electron-proton system, most of the radiation

me re  2e  Fe  e 2 /(re  rp )2 . But with   2 e , the effective total acceleration is a  ae  2 2 .


The total radiation power is then

Ptotal radiated  2 4
2 e2 2
3c 3

ae  2 5 e6 / me2  3c 3 (re  rp )4 . (11)

due to the radiation. The balance requires Ptorquing  Ptotal radiated , or


Now posit a balance between the energy gain rate due to the torque and the energy loss rate

( e 4 / mp ) c(re  rp )3  (2 5 e 6 / me2 ) 3c 3 (re  rp )4 . (12)

This equation can be solved for

re  rp  32 mp e 2 3me2c 2  5.5  10 9 . (13)

Compare that value to the accepted value re  rp  5.28  10 9 cm. The match is fairly close,
running just about 4% high. That means the concept of torque vs. radiation does a fairly
good job predicting the ground state of Hydrogen.

In conventional QM, re  rp is expressed in terms of Planck’s constant h , which is


4.4 Unification of physics via Planck’s constant

presumed to be a fundamental constant of Nature:

re  rp  h 2 4 2  e 2 . (14)

Here  is the so-called ‘reduced mass’, defined by  1  me 1  mp 1 , which makes   me .


Using that approximation and equating the two expressions (13) and (14) for re  rp implies

 e2
h 128mp / 3me . (15)
c

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 139

This expression comes to a value of 6.77  10 34 Joule-sec, about 2% high compared to the
accepted value of 6.626176  10 34 Joule-sec. This reasonable degree of closeness suggests
that Planck’s constant may reasonably be considered a possible function of other
fundamental constants of Nature, and so not itself an independent fundamental constant of
Nature. Or the situation may reasonably be considered the other way around: that some
other fundamental constant of Nature is really a function of Planck’s constant. Either way,
we would have one less independent fundamental constant of Nature, and that would mean
one more degree of unification among the different branches of physics.
But of course, the expression for h developed here can fulfill such aspirations only if the
theory being developed can do a great deal more than just match the ground state of
Hydrogen. Worthy targets for additional work include: anticipating the story for isotopes of
Hydrogen, anticipating from there what happens with other elements, explaining the
excited states of Hydrogen and their resulting spectral lines, anticipating from there the
spectral features of some other elements, and characterizing the behavior of the full database
on ionization potentials of all elements, and much more. It all constitutes a developing
research area that I refer to as ‘Algebraic Chemistry’.

5. Extensions and extrapolations from hydrogen


5.1 Larger nuclear mass
The negative energy of the electron in the ground state of the Hydrogen is

e 2 (re  rp )  3c 2 me2 2 5 mp . (16)

This is the energy that would have to be provided to liberate the electron, or ionize the atom:
the ‘ionization potential’.
Eq. (16) provides the basis from which to build corresponding expressions for other entities.
For example, the extension to Deuterium and/or Tritium requires that the proton mass mp
be replaced with a more generic nuclear mass M , and that rp be replaced by rM . Then we
have for the ionization potential of this more massive system:

e 2 (re  rM )  3c 2 me2 2 5 M . (17)

5.2 Arbitrary nuclear charge


The extension of the model to a neutral atom with nuclear charge number Z involves Z
electrons as well. To develop the mathematical model, we must return to the expressions for
Ptorquing and Ptotal radiated , Eqs. (10) and (11). All the factors of e 2 change to Z 2 e 2 , and the

factor of me2 changes to Z 2 me2 . The equality Ptorquing  Ptotal radiated becomes

Z Ptorquing  (Z / Z )Ptotal radiated  Z Ptotal radiated . So nothing happens to the equality


4 6 2 4

between Ptorquing and Ptotal radiated , Eq. (12). But for the more charged system, the energy Eq.
(17) becomes

Z 2 e 2 ( re  rp )  Z 2 3c 2 me2 2 5 M . (18)

www.intechopen.com
140 Theoretical Concepts of Quantum Mechanics

This scaled-up expression represents the magnitude of the total ionization potential of the
system involving Z protons and Z electrons. What is then comparable to the ionization
potential for removing a single electron is:

   
Z e 2 (re  rM )  Z  3c 2 me2 2 5 M  (Z / M )  3c 2 me2 2 5 . (19)

Thus in the math we find a Z / M scaling law. What do we find in the actual data?
Something much more complicated, and indeed so complicated that we would be unlikely
ever to figure it out without the clue that Z / M is part of the story. The involvement of M
means the involvement of isotopes, and unwanted complexity. So the clue tells us to look at
ionization potentials, not in raw form, but scaled by M / Z , to remove the Z / M factor that
the math anticipates.
Figure 2 shows the pattern found. Seven orders of ionization are included. There is a
fascinating, but lengthy, story about ionization orders 2 and up; see Whitney (2012). The
part of it that will be most important for the present development is obvious from Fig. 2: the
energy required to completely strip the atom scales with Z 2 .

10000

1000

100

10

1
0 10 20 30 40 50 60 70 80 90 100 110 120

Fig. 2. Ionization potentials, scaled by M / Z and modeled algebraically.

equal to that of Hydrogen, IP1,1 , and an increment IP1,Z . The increment arises from
With their M / Z scaling, all of the IP ’s can be represented in terms of a baseline value

interactions just between the electrons, quite apart from the nucleus. The electron-on-
electron increments are very regular in their behavior. First of all, every period exhibits a

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 141

general rise, and by the same factor of 7 / 2 . Second, there is a general drop from one
period to the next, for the first three periods, and all by the same factor of 7 / 8 .
Then within periods, there is a very regular pattern. There are sub-period rises keyed to the
traditional ‘angular momentum’ quantum number l , and to a non-traditional parameter N

period as 2 N 2 . For l  0 , we have:


that goes 1, 2, 2, 3, 3, 4, 4 for periods 1 through 7 , and gives the number of elements in a

incremental rise  total rise  fraction , and fraction  (2l  1) / N 2  ( N  l ) / l  . (20)

The following Table details the behavior fractional rises in First-order IP ’s over all sub-
periods:

period: N : l : fraction: l : fraction: l : fraction: l : fraction:


1 1 0 1
2 2 0 1/2 1 3/4
3 2 0 1/3 1 3/4
4 3 0 1/4 2 5 / 18 1 2 /3
5 3 0 1/4 2 5 / 18 1 2 /3
6 4 0 1/4 3 7 / 48 2 5 / 16 1 9 / 16
7 4 0 1/4 3 7 / 48 2 5 / 16 1 9 / 16

The scaled ionization potentials are called IP ’s. They are meant to be ‘population generic’;
that is, the information they contain concerning one element can be applied to a calculation
about another element in a different state of ionization, or excitation, by applying the Z / M
appropriate for the second element and its state.

5.3 Unequal counts for electrons and protons


Let us first consider ionization sates. These are important for applications in Chemistry,
since chemical reactions involve ions. With all this regularity displayed in Fig. 1, it should be
possible to use it to help predict the energy budget for all sorts of chemical reactions. We
just need a rational way to extrapolate from all the formulae representing the regularities for
single electrons being removed from neutral atoms to formulae for electrons being removed
from, or added to, ions of all sorts.
Generally, if an atom is in an ionized state, then in place of just Z we have an electron count
Ze distinct from the proton count Zp . The electron-on-electron interaction does not involve
the nucleus, and so always scales with Ze / M . But electron-nucleus interaction previously
represented by (Z / M )IP1,1 now has to involve both Ze and Zp . We have for the total

  
system


ZpZe e 2 (re  rM )  ZpZe M  3c 2 me2 2 5 . (21)

What is then generally comparable to the nuclear-orbit part of the ionization potential for
removing a single electron? To develop an answer to this question, we must return again to
the expressions for Ptorquing and Ptotal radiated , Eqs. (10) and (11). Clearly, all of the factors of

www.intechopen.com
142 Theoretical Concepts of Quantum Mechanics

e 2 change to ZpZe e 2 . It is as if all factors of e changed to ZpZe e . Removal of one electron


is then like removal of one ZpZe e charge. What is comparable to the ionization potential

   
for removing a single electron from the ion is then

ZpZe e 2 (re  rM )  ZpZe M  3c 2 me2 2 5 . (22)

Thus for ions, we see in the math a ZpZe M (Zp ) scaling law for that part of the ionization
potential that reflects electron-nucleus interaction, IP1,1 . So for computations we use:

IP1,1Z / M  IP1,1 ZpZe M(Zp ) . (23)

interactions, IP1,Z , the relevant Z is Ze . But the relevant M is still M(Zp ) , the only
For the other part of the ionization potential, that reflecting just the electron-on-electron

significant mass in the problem. So for computations we use:

IP1,Z Z / M  IP1,Ze Ze M(Zp ) . (24)

This basic information can help one to model the energy budget for any chemical reaction.
To assist readers who want to try this out, the necessary data displayed by Fig. 1 is tabulated

Here is one small example. Recall the comment about Fig. 1 that, for nuclear charge Z  2
in numerical form as Appendix 1 at the end of this Chapter.

and up, the energy required to completely strip the atom scales with Z2 . The actual formula
plotted on Fig. 1 goes

IPZ , Z  2 IP1,1  Z 2  2  14.250  Z 2 . (25)

The resulting IPZ , Z is population-generic. The corresponding element-specific quantity is

total stripping is 2  14.250  Z 3 / M(Zp ) eV’s.


IPZ ,Z multiplied by the factor Z / M( Zp ) . Thus the element-specific energy requirement for

We can now compare the total energy required to strip an atom one electron at a time with
the energy required to strip it of its electrons all at once. The two elements Helium and
Lithium are good examples because they represent the extremes of very high first-order
ionization potential and very low first-order ionization potential. The data for them in
numerical form comes from Appendix 1. Here is how the calculations go:

Helium: ( Zp  2 , M(Zp )  M 2  4.003 )

Write Formulae:

2 He  2 He  : IP1,1  2 / M2  IP1,2  2 / M2 ; 2 He+  2 He  : IP1,1  2  1 M2 .

Insert Data:

2 He  2 He  : 14.250  2 / 4.003  35.625  2 / 4.003 ; 2 He+  2 He  : 14.250  1.4142 4.003 .

Evaluate Formulae:

2 He  2 He  : 7.1197  17.7992  24.9189 ; 2 He+  2 He  : 5.0343 .

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 143

Evaluate Total Stripping One-at-a-Time:

2 He  2 He  : 24.9189  5.0343  29.9532 eV’s.

Compare to Total Stripping All-at-Once:

2  14.250  Z 3 / M(Zp )  2  14.250  2 3 / M2  2  14.250  2 3 / 4.003  56.957 eV’s.

Lithium: ( Zp  3 , M(Zp )  M 3  6.941 )

Write Formulae:

3 Li  3 Li + : IP1,1  3 / M 3  IP1,3  3 / M3  IP1,2  2 / M3 ;

3 Li
+
 3 Li ++ : IP1,1  3  2 M3  IP1,2  2 / M3 ; 3 Li ++  3 Li 3+ : IP1,1  3  1 M 3 .

Insert Data:

3 Li  3 Li + : 14.250  3 / 6.941  ( 1.781)  3 / 6.941  35.625  2 / 6.941 ;

3 Li
+
 3 Li ++ : 14.250  2.449 6.941  35.625  2 / 6.941 ; 3 Li ++  3 Li 3+ : 14.250  1.7321 6.941 .

Evaluate Formulae:

3 Li  3 Li + : 6.1591  0.7698  10.2651  4.8758 ;

3 Li
+
 3 Li ++ : 5.0278  10.2651  15.2929 ; 3 Li ++  3 Li 3+ : 3.5560 .

Evaluate Total Stripping One-at-a-Time:

3 Li  3 Li 3+ : 4.8758  15.2929  3.5560  13.9731 eV’s.

Compare to Total Stripping All-at-Once:

2  14.250  33 / M(Z3 )  2  14.250  33 / M3  2  14.250  33 / 6.941  110.8630 eV’s.

In these two examples, we see that removal of all the electrons, all at once, takes much more
energy than removing the electrons one electron at a time. It is plain to see that total stripping
all-at-once is a vigorous, even violent, event. It is the stuff of special-purpose laboratory or
field investigation. By contrast, total stripping one-at-a-time is a gentle process. The one-at-a-
time process is an example of the stuff of ordinary production Chemistry.

5.4 Excited states - hydrogen


Now let us begin to consider excitation states. These are key for understanding emission or
absorption spectra, a fabulously rich source of data about atoms. But atomic spectra are
complicated. The standard way to begin to understand them is mathematically, from the
family of solutions provided by the differential equation that Schrödinger postulated for the
abstract wave function characterizing the electron in the Hydrogen atom.

www.intechopen.com
144 Theoretical Concepts of Quantum Mechanics

negative energy, E , determined largely by a principle quantum number n  1, 2, 3...


The standard QM view is that the Hydrogen atom has multiple ‘stable states’, each with

according to En  E1 / n2 . The idea is that the electron can reside in an upper state
( n  1 ), but only rather precariously, and when it teeters and falls back to the ground state
( n  1 ), a photon is emitted.
But the Hydrogen atom has only two constituent particles, the electron and the proton, and
thus very few classical degrees of freedom. That fact makes it difficult to imagine an infinite
multiplicity of different ‘states’ that a Hydrogen atom could exhibit. We are left to ponder a
mystery of mathematical QM. So it is tempting to try to develop an additional, more
immediately physical, way of understanding the spectral complexity that we see. Consider
the possibility that individual Hydrogen atoms may not, by themselves, actually have
excited states. Instead, the term ‘excited state’ may be better applied to a system that
involves several Hydrogen atoms.
Key to this idea is that charges can form entities called ‘charge clusters’. [Concerning charge
clusters at the macro scale of laboratory experiments and field observations: see, for
example, Beckmann (1990), Aspden (1990), Piestrup and Puthoff (1998).]
Evidence concerning the probable existence of charge clusters at the micro scale of atoms is
plainly visible in the data on IP ’s (Fig. 1): some electron counts are very stable and hard to
break apart (e.g. noble gasses), while some electron counts are very un-stable and hard to
keep together (e.g. alkali metals). Why would electron counts matter so much if the electrons
were not in deep relationships with each other?
But how can electrons outwit electrostatic repulsion? Once given the clue that they evidently
can do this, it becomes possible to imagine how they might do it. The key is that electrostatic
repulsion dominates in a static situation. In a dynamic situation, electrons may move at
speeds exceeding light speed (Remember, Sect. 3 cast doubt on the founding postulate of
SRT, and SRT is all there is to forbid superluminal speeds.). If so, a repulsion signal from
one electron may reach another electron only by the time the first electron has moved so
much that the repulsion from its ‘then’ position has become the attraction to its ‘now’
position. In fact, multiple electrons can form circulating ring structures that are quite stable
(for details, see Whitney 2012).
So consider the possibility that an excited state of Hydrogen is actually nH neutral
Hydrogen atoms, with the nH electrons in a negative charge cluster, and the nH protons in
a positive charge cluster, making a kind of ‘super’ Hydrogen atom; i.e., Hydrogen with
every factor of electron mass me , proton mass mp and charge e scaled by nH . The
torquing power PT  ( e 4 / mpc ) (re  rp )3 then scales by (nH )4 nH  ( nH )3 , and the radiation
power PR  2 5 e6 3c 3 (me )2 (re  rp )4 scales by (nH )6 (nH )2  (nH )4 . The solution for system
radius re  rp  32 mp e 2 / 3(me )2 c 2 then becomes:

rnH e  rnH p  32(nHmp )(nH e )2 / 3(nH me )2 c 2  nH (re  rp ) . (26)

i.e., the system radius is scales with nH . The system orbital energy

E1   2   2 e 2  3(me )2 c 2 32 mp e 2 becomes
2
e
(re  rp )
1 1

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 145

EnH   2 nH   2 nH  e 2  3(me )2 c 2 32 mp e 2  .
1 (n )
2
e2
H ( re  rp )
1
(27)

i.e., the system orbital energy also scales with nH . This result is the same as if the atoms
were isolated, instead of being organized into a big system with two charge clusters. This
suggests that the energy available for generating photons by de-excitation isn’t ‘orbital’ at
all, but is instead the energy tied up in forming the charge clusters out of the multiple
electrons and the multiple protons from the multiple Hydrogen atoms.
What can we infer about such charge clusters? As in the modeling of IP ’s for ions, we can
again consider Fig. 1 as a source of information about electron clusters of sizes up to 118,

clear that most of the IP ’s are positive, meaning their electron clusters are hard to break.
quite apart from the particular element that the information is located with. From Fig. 1, It is

So despite being made of same-sign charges, most of them exist in negative energy states.

Z  2, 10, 18, 36, 54, 86, (118) . These elements are at the ends of periods on the periodic table,
The ones that are particularly hard to break are the ones associated with the noble gasses:

and the lengths of the periods themselves are: 2, 8, 8 , 18, 18, 32, (32) . (Parentheses mean we
haven’t discovered, or created, that element yet.) The implication is that excited states of

nH  2, 8 , 18, 32, ...
Hydrogen existing in the form of ‘super Hydrogen’ would most frequently exist with

started with nH  32 . It could, for example, decompose into 18, 8, 2, 2, and 1, 1; i.e. some less
Can we anticipate what would happen when any such excited state de-excites? Suppose we

excited states and a couple of ground-state atoms; 6 daughter systems in all. Suppose that
for every such daughter produced, there is a photon released. Exactly how might that work?
Observe that four daughters are in states that are even more negative than the starting state,
so those are no problem. But two daughters are in the ground states, which is not more
negative than the starting state. So energy from the other daughters has to be enlisted to
create any photons there.
For any nH , there may be a de-excitation path, or paths, for which the energy budget is
insufficient, in which case those paths won’t be taken. There may also be de-excitation paths
for which the energy budget is more than sufficient, in which case there will be, not only
spectral radiation, but also a bit of heat radiation. Very rarely, there might be a de-excitation
path for which the energy budget is just exactly right.
The spectral lines that occur with Hydrogen (or any element) are typically characterized in
part by differences in inverse square integers. The integers involved are traditionally
understood in terms of the familiar radial quantum number n . Is it possible to understand
them also in terms of the nH used here?
Recall that if one then chooses to model the behavior of Hydrogen ‘excitation’ in terms of a

n , then the orbit-radius scaling has to be the quadratic scaling r1  rn  n2 r1 of standard


single Hydrogen atom with discrete radial states identified with the radial quantum number

QM, not the linear orbit-radius scaling re  rp  rnH  nH (re  rp ) of the present model. So
why does one way of looking at the problem involve a quadratic n2 , while the other way of
looking at it involves a linear nH ?

nH  2, 8 , 18, 32,... corresponding to the lengths of the rows in the Periodic Table. These row
Recall that there was good reason to suggest highest probability for the values

lengths can be characterized as 2N 2 for N  1, 2 ,  2 , 3 , 3 , 4 , (4) . So nH actually does encode
something that is quadratic, namely the N 2 , and is therefore similar to the quadratic n2 .

www.intechopen.com
146 Theoretical Concepts of Quantum Mechanics

5.5 Beyond both hydrogen and ground states: Spectroscopy


In spectroscopy, we observe light created when an atomic system relaxes in some way. For
elements beyond Hydrogen, the spectral lines that occur are often characterized in part by
the so-called Rydberg factor:

2 2 me e 4
R 
Z2
ch 3 1  me / M
. (28)

The R is traditionally interpreted as the energy needed for total removal of one electron

from a state labeled n1 to a higher state labeled n2  n1 , and conversely the energy released
from the ground state to infinity, leaving an ion. The energy needed for an electron to get

when it goes back to n1 , is then modeled as

E  R  1 /( n1 )2  1 /( n2 )2  . (29)

Observe that R contains a factor of Z 2 , just like the IP ’s for total ionization, IPZ ,Z of Eq.
(25) do. That means R is referring to the absolutely largest photon energy that the system
could ever possibly be imagined deliver: starting from a state of total ionization, i.e. a naked
nucleus, and having the entire electron population return in one fell swoop, with the
emission of just one photon for the whole job. That scenario could never actually happen.

integers in the square bracket bring E down to values appropriate for one-at-a-time
One-at-a-time electron return is the only plausible return scenario. The inverse square

scenarios.
Observe that the Rydberg model for spectral lines already conflicts with an older model for
the atom developed from the PT; i.e., electron ‘shells’ enclosing the nucleus, inner shells
filled, and at most one outer shell unfilled; partially filled for most elements, and completely
filled for noble gasses. The older PT-based model suggests shielding of the nucleus by the
filled inner shells of electrons. But the occurrence of a Z 2 in R , even for large n1 and n2
in E , means there is no shielding of the nucleus. So electrons must be in tight clusters,
rather than nucleus-enclosing shells.
Observe that R does not contain any Z / M factor like the IP model contains. Instead, it
has a factor

1 (1  me / M )  M /( M  me ) , (30)

which is essentially unity. At the time when R was formulated, most of the known trans-
Hydrogenic elements had M  2Z , and the factor of Z / M  1 / 2 could be absorbed into
an external constant factor, the 2 2 me e 4 ch 3 in R . That is no longer the case today. We
know about heavy elements for which M  2.5Z , or Z / M  2 / 5 . So now it would be
better to use the function Z / M instead of the number 1 / 2 . An extra bonus would be that
Hydrogen, with Z / M  1 , would be included.
Now consider that spectral lines might not to arise from de-ionizing one ion of one atom, but
rather from de-exciting a system involving multiple neutral atoms. In this description, the
n1 and n2 are not identifiers of different states of one atom, but rather numbers of atoms
organized into super atoms. Otherwise, nothing really changes. However we interpret their
meaning, the predicted spectral lines remain the same.

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 147

6. Unification between Newton and Maxwell


This last technical Section of this Chapter returns to the first physics disunity mentioned in
the Introduction: the seemingly different coordinate-transformation properties of Newton’s
Laws for mechanics and Maxwell’s equations for electrodynamics. Newton’s laws are form
invariant under Galilean transformations. But Maxwell’s equations are generally thought to
be form invariant only under Lorentz transformations. Especially, they are thought to be not
form invariant under Galilean transformations.
So a curious situation exists within physics today. It is generally expected that the equations
of physics should be tensor equations. By definition of the word ‘tensor’, a tensor equation is
form invariant under arbitrary changes of reference frame, assuming no singularities or
other cruel and unusual circumstances in the transformation or its inverse transformation.
That means a tensor equation should be form invariant under arbitrary, though reasonably
well-behaved, space-time transformations.
So, are Maxwell’s equations really tensor equations? Or not? Mathematicians have good
reason to challenge the believed tensor status of Maxwell’s equations, while physicists have
good reason to challenge the believed requirement for invariance under anything other than
Lorentz transformation. But the situation is not generally acknowledged. It is the proverbial
‘elephant in the living room’.
Clarifying this situation can assist physics in becoming more unified from its beginning to
its present. And mathematics has lots of applicable tools; see Kiein (2009). The present work
offers an approach that is also mathematical, but a lot more elementary. Maybe it will
communicate to different readers.
The problem, I believe, is of a type with which QM has some history. QM appears to be the
first branch of physics that well and truly needed complex numbers. They may have been
used in physics before QM, but they were only one of the tools available for the problems
then at hand. Sines and cosines could generally handle any problem just as well as complex
exponentials could handle it. But with QM, complex exponentials became truly essential for
doing physics.
The history of mathematics has been a tale of increasing range of objects included in the
discussion. It began with real, positive integers; it grew with the inclusion of zero and
negative integers, and grew again with the inclusion of all rational numbers, and again with
the inclusion of all irrational numbers. Then it grew with the inclusion of imaginary
numbers, thus creating complex numbers. This was the first of a number of ‘doublings’ of
the number of dimensions attributed to mathematical objects. [See Rowlands (2007).] After
complex numbers, we got quaternions, and bi-quaternions, or octonians, and there is no
reason to suppose that further doublings will not continue to prove useful.

for example, the square root of 3 . It cannot be evaluated within the real number system,
Complex numbers make possible operations that are not possible without them. Consider,

but in the complex number system, it is just  i 3 .


I believe ‘doublings’ are generally like this: they make possible operations that were not
possible without them. There appears to be today an opportunity for a doubling in the realm
of tensor calculus. There are presently exactly two tensor-transformation behaviors
identified, called ‘contravariant’ and ‘covariant’. It appears that tensor calculus can be
usefully extended through a doubling of the number of transformation behaviors that can be
described, from two to four. It appears that such a doubling can resolve the apparent

www.intechopen.com
148 Theoretical Concepts of Quantum Mechanics

conflict between Newtonian and Maxwellian physics: it can make possible a display
showing how Maxwell’s equations can actually be form invariant under arbitrary
coordinate transformations.

6.1 The opportunity offered by tensor notation


The display of four transformation behaviors requires the use of four tensor index
positions. So in addition to the usual contravariant (index up-right) and covariant (index
down-right) positions on the right side of a tensor symbol, we need to us the positions
available on the left side: index up-left and index down-left. Since left-side index positions
have not been in used in this new way before, they need new names designed for the
purpose. To recall the move from right to left, let us use the prefix ‘trans’. So let the up-left
index position be called ‘transcontravariant’, and let the down-left index position be
called ‘transcovariant’).
All the transformations are describing what happens to tensor merates when the frame of
reference changes; i.e. when the basis unit vectors defining the frame of references are
replaced with other basis unit vectors. The transformations discussed here are arbitrary
within the specifications that make the connections between reference frames reasonably
well behaved; the individual relationships are differentiable and reversible, the matrix
representations of them are invertible and unimodular.
I mention both tensors and matrices because they are equivalent notation schemes that
can be used interchangeably for describing systems of linear equations. Tensor notation is
useful for making a compact statement of a whole mathematical situation. Matrix notation
is useful for separating a whole mathematical situation into constituent parts for
calculations. Individual linear equations are useful for focusing on individual parts of the
mathematical problem. Human beings do have strong personal preferences about which
approach to use, but all of these approaches should agree on the basic facts of a given
situation, so any of these approaches should be acceptable. In the present work, all
approaches will be used. That way, everyone can find something to like, and everyone can
find something to dislike!
In the case of the matrix displays and the linear equations, the presentation does save a little
space by ignoring two spatial dimensions and focusing on one spatial dimension (call it 1)
and the temporal dimension (call it 0).

6.2 Transformation of a contravariant object


The most familiar transformation is the contravariant one. The prefix ‘contra’ means these
tensor merates change opposite to the way the basis unit vectors of the reference frame
change. For an arbitrary input vector X , the transformation reads X   x  / x  X ,
where we see the transformation as partial derivatives of coordinates, new with respect to
old. Equivalently X   T X , where we see the transformation written as the tensor T .
  X 0     T00    T10      X 0  
Also equivalently, we have     , where we see everything, the input and
  X 1     T01     T11     X 1  
output vectors and the transformation, in matrix format. Or equivalently, we have
X 0  T00 X 0  T10 X 1 and X 1  T01X 0  T11X 1 as two separate linear equations.

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 149

  T00     T10  
 1 
  T0     T11  
For the contravariant transformation matrix one can define a reverse

  R0    R 0  
transformation matrix  0 1  , wherein R00  T00 and R11  T11 , but R10  T10 and
  R01    R11  
R01  T01 . Applied to X  , the reverse transformation R takes X  back to X :
X   R X  . That is to say: X 0  R00 X 0  R10 X 1 and X 1  R01X 0  R11X 1 . Expressed in matrix
  X 0     R00    R10     X 0  
form, the reverse transformation is the inverse transformation:     1 1   1  , or
  X     R0    R1     X  
1

  R00    R10     T00    T10     1   0  


    . The distinction between the words reverse and inverse is nil in
  R01    R11     T01    T11     0    1 
the contravariant context. But it becomes important in the next context.

6.3 Transformation of a covariant object


The prefix ‘contra’ means reverse to the prefix ‘co’. The covariant transformation goes the
same way the basis unit vectors change. So the covariant transformation X   C  X in
  X     C 0    C 1     X  
matrix format  0    0 0   0  uses transformation matrix C equal to the reverse
  X1     C 1    C 1     X1  
0 1

  X0     R00    R10     X0 


R:    ,
  X1     R01    R11     X1 
contravariant transformation matrix or equivalently

X0  R00 X0  R10 X1 and X1  R01X0  R11X1 . It is generally assumed that this is the same as
saying the covariant transformation is the inverse to the contravariant transformation.
Notice however that the off-diagonal merates C 01  R10 , and C 10  R01 have indices switched
around. This is because C operates on a covariant object, whereas, in its original definition,
R operated on a contravariant object.
The index switching makes no difference if we limit attention to transformations that are
space-time symmetric, i.e. Lorentz transformations. But if we wish to investigate any other
type of transformation, we have to investigate whether the switch makes a difference.
Consider the inner product X  X  . Under Lorentz transformation, it is preserved, equal to
X  X   X X . But if we do not have space-time symmetry, is it still preserved? This
question has to be answered by testing.
Laying out the problem in matrix format, we have to make one of the vectors, say the
covariant one, a row vector, and then we have to test:

  X 0     R 0    R 1     T 0    T 0     X 0     X 0  
  X0    X1        X 0    X1    0 0   0 1     ?    X0    X1    
  X     R1    R1     T0    T1     X     X  
(31)
1 0 1 1 1 1 1

www.intechopen.com
150 Theoretical Concepts of Quantum Mechanics

Observe that the R matrix is transposed from what it would need to be to make the RT
matrix product collapse to the identity. So the inner product X X is generally not
preserved if we do not have space-time symmetry.

6.4 Transformations for objects of four types


In order to recover the general availability of preserved inner products, the two additional
transformation behaviors are defined. The transcovariant transformation is defined as the
transposed inverse of the contravariant one. The transcontravariant transformation is
defined as the transposed inverse of the covariant one.
Recall that this discussion began with the contravariant transformation written in the tensor
notation X   x  / x  X . The discussion soon became complicated enough to merit
introduction of more detailed notation that can clearly distinguish the four cases. The
following Table illustrates the expanded tensor notation:

 
transcontravariant,  contravariant,

X    T   X  x  / x   
X X   T  X  x  / x  X
reverse direct

transcovariant, covariant,
    
 X    C   X x / x  X X   C   X  x  / x   X
direct reverse

The Table is organized for user convenience, with the position of information corresponding
to the index position: upper right for contravariant, lower right for covariant, lower left for
transcovariant, and upper left for transcontravariant. The index position assigned to an
object determines the transformation law that it follows.
Now let two arbitrary numbers with magnitude less than unity be represented by the letters
A and B (chosen from the word ‘arbitrary’!). Let the arbitrary numbers represent in turn
the off-diagonal elements of transformation matrices. The following table shows the
corresponding matrix notation:

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 151

transcontravariant, contravariant,
  0 X      1    A     X     X 0      1    B    X  
1     B    1   1   1    A    1    1 
0 0
1 1
   X   1  AB      X     X   1  AB     X  

transcovariant, covariant,
  0 X      1    A    0 X     X0      1    B    X0  
     B   1      X        A   1    X  
1 1
 1 
  X   1  AB   1   1 
 X   1  AB   1 

Observe that this Table uses negative signs on the arbitrary A and B in the contravariant

sign choice is used to help recall the prefixes ‘contra’ and ‘co’. Observe too that if B  A , we
and transcontravariant cases, positive signs in the covariant and transcovariant cases. This

finally that if B  0 , we have universal time, which is the case of Galilean transformations.
have space-time symmetry, which is the case of Lorentz transformations. And observe

But A and B are arbitrary, and so can also represent other transformations as yet
unnamed.

6.5 Transformations for invariant objects


The underlying purpose of tensor calculus is to focus on mathematical objects that are
‘coordinate free’, or ‘frame independent’, or ‘invariant’ (whether in form or in numerical
value), – all expressions meaning that coordinate transformation does not change anything
fundamental about an object so-described: values of scalars, or relationships expressed as
equations involving tensors.
The user of tensor calculus expects certain behaviors. There should be number invariant
inner products of vectors and of higher-order tensors. The ‘unity’, or ‘Kronecker delta’ is not
presently regarded as a real tensor, but can be accepted as one if it can be demonstrated
number invariant. Finally, the user will certainly expect a number invariant ‘metric tensor’,
the essential tool for manipulating index positions to develop tensor equations. Displaying
that all these expectations can be met in the case of arbitrary transformations, not just
Lorentz transformations, is the objective of this Sub-Section.
The matrix notation is useful in checking out the transformation of all these entities. For
example, the preserved inner product of a vector X with itself looks like (note the
transpositions for operating on row vectors):

  X 0   1    1     B      1     B    X     X 0  


XX   0 X  1 X       0 X    1 X          A    1   1      1
0
 

  X   1  AB    A    1       X     X  



 0
X    1
X    0
X    1
X   
XX . (32)
1

www.intechopen.com
152 Theoretical Concepts of Quantum Mechanics

or

  X 0   1    1     B    1     B    X 0     X 0   


XX      X     X         X     X             X     X      XX . (33)

  X 1   1  AB    A    1      A    1     X 1     X 1  
0 1 0 1 0 1

The more familiar inner product X X is preserved with Lorentz transformations, but not
with arbitrary transformations. So it shouldn’t be considered any kind of ‘invariant’. The
same is true of the unfamiliar (  X )(  X ) .
With the extended tensor notation, we can identify the index positions that definitely make
a number invariant Kronecker delta. It looks like (note the transpositions for operating on
row vectors):

  1   1    A    1   0     1    A     1   0   


   
1  AB    B   1    0    1 
   B   1     0    1     .
   
(34)

or

1   1    A     1   0     1    A    1   0   



          .
1  AB    B   1    0    1     B   1    0    1 
(35)

The more familiar   is preserved with Lorentz transformations, but not with arbitrary
transformations. That is why it does not qualify as a tensor. The same is true of the


unfamiliar .
Some readers will be surprised to see the present argument using the Lorentz metric,
  1    0   
  0    1  , without accepting a limitation to Loentz transformations. It is widely supposed
 
that the Lorentz metric requires Lorentz transformations, and/or Lorentz transformations
require the Lorentz metric. But such a connection is not in fact mandatory.
The generally preserved forms of the Lorentz metric tensor look like (note the transpositions
for operating on row vectors):

1    1    A    1    0       1    A  1   1    A     1    A    1    0     

g     B   1     0    1     B   1          g . (36)
1  AB      1  AB    B   1     B   1   0    1
and

1   1    A    1    0      1    A  1   1    A     1    A     1    0   


g        B   1        B   1     0    1    g . (37)
1  AB    B   1    0    1   AB    B   1 

  1    

The more familiar g  and g  are preserved with Lorentz transformations, but not with
arbitrary transformations. They shouldn’t be considered any kind of ‘invariant’. The same is

true of the unfamiliar g and  g.

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 153


The number invariant g  and  g  can function to raise and lower indices on objects. For
example, X   (  g )X 
and X  (  g  ) X , or  X  (  g  ) (  X ) and a X  ( a gb ) ( b X ) .

One can also write additional index assignments for g . Altogether, there are 10 possible
assignments, as there are 4  3 / 2  6 with indices in different corners, and 4 with indices in
the same corner.
Two of the additional index assignments look like  g and  g . These two entities cannot
do anything to an index except change its name. For example, X   X  (  g ) and
X  X  (  g ) , or X  (  g ) (  X ) and 
X  (  g ) (  X ) . So  g and 
g are just
equivalent to the number invariant   
and  already noted above.

Further additional index assignments on g create entities that can serve to convert a regular
index into a trans one, or a trans one into a regular one. None of these entities are number
invariant, but in practice, that does not matter. The user does not convert just a single object;
the user converts a whole tensor equation. The index-converting g entities typically occur
in pairs, and the pairs contract to number invariant objects. When they don’t occur in pairs,
they do occur on both sides of an equation, and cannot affect the issue of equation form
invariance.
Another two of these of g ’s are g and  g . They function to do  X  (  g )X and

X  (  g )X  , or X   g (  X ) and X   g (  X ) . As a pair, they contract to


(  g )( g )   g , or to (  g )( g )   g , both of which are number invariant.

The final four indexed g ’s are g,  g , g , and g . They can all function to change a
regular index into a trans one, or a trans one into a regular one, but with a twist: ‘co’ goes to
‘contra’, or ‘contra’ goes to ‘co’. That is,  X   (  g ) X  ,  X  (  g ) X  , g (  X )  X  and
g (  X )  X  . As noted above, the contractions (  g )( g )   g and (  g )( g )   g are
number invariant.
The bottom line is this: to be sure of invariance under arbitrary transformation, not just
Lorentz transformation, always contract a regular index with a trans index.

6.6 General invariance for Maxwell’s equations


Maxwell’s equations in current tensor notation read:

4 
 F  J and  D  0  . (38)
c

The two-index F and D tensors refer to the electromagnetic field and the ‘dual’
thereof. The electromagnetic field tensor F has merates that are components of the three-
dimensional electric and magnetic field vectors, E and B . The D is the dual to F ,
whose merates are components of B and E . The one-index tensors J  and  refer to the
source charge-current density vector and the differential operator vector. The indices  and
 take the four values 0,1, 2, 3 .

www.intechopen.com
154 Theoretical Concepts of Quantum Mechanics

The seeming limitation of Maxwell’s equations to invariance only under Lorentz


transformation arises entirely from the differential operator being written as a covariant
vector. In the extended tensor algebra, this operator is identified as transcovariant, and then
Maxwell’s equations look like:

4 
(   )F  J and (   )D  0  . (39)
c
Written this way, Maxwell’s equations are manifestly form invariant, not only under
Lorentz transformation, but also under any arbitrary (just well-behaved) transformation,
including Galilean transformation.

7. Conclusions
About Maxwell’s equations and photons: Photons have a life history that begins with
emission as an electromagnetic pulse pulse, proceeds with development into a waveform,
then changes into regression back to a pulse, and ends with absorption by a receiver. This
life history of the photon can be modeled by imagining some mirrors that apply boundary
conditions corresponding to the desired scenario, feeding a Gaussian pulse at the source to
Maxwell’s equations, watching Hermite polynomials then emerge, and then finally pile up
at the receiver.
About EM signals and photons: The life history of the photon suggests that the
assumption upon which Einstein’s SRT is founded is over-simplified. If we will make the
founding assumption more realistic, then we will get more believable results. The more
believable results can help us reconcile SRT with the QM of atoms. We can understand
why Planck’s constant occurs. It represents the balance between competing phenomena:
on the one hand, energy loss due to radiation from accelerating charges; on the other
hand, energy gain due to internal torquing within the atomic system due to finite speed of
signal propagation.
About Atoms: Viewed in the right way, chemical and spectroscopic data reveal a
tremendous amount of regularity. So we are well enabled to interpolate and extrapolate for
situations where actual data is not available. We can analyze scenarios where electrons are
subtracted from or added to an atom, all at once, or one at a time; whatever we need. But
take care: in the existing literature, the distinction between ‘all-at-once’ and ‘one-at-a-time’ is
often obscure, so be careful.
About Maxwell and Newton: There should have been no conflict between Maxwell’s
equations and Newton’s equations over the issue of transformation invariance. Maxwell’s
equations are form invariant under Galilean transformations, just as they are form invariant
under Lorentz transformations. Physics does not have conflicts. Only people have conflicts.
And people can resolve their conflicts. The conflict perceived in the case of Newton vs.
Maxwell is resolved with an extension of mathematical formalism.
About Physics in General: This work has shown that SRT deserves a moment of caution,
and the reader may reasonably worry that GRT deserves some caution too. So it may be
premature to develop a theory of quantum gravity. Placing the QG capstone onto the RT
and QM pillars of 20th century physics may produce something that resembles the ancient
constructions at Stonehenge, but not the Gothic cathedrals of Europe, much less anything
modern.

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 155

8. Appendix 1. Numerical data on ionization potentials for all elements

Charge Mass Ionization IP  Ionization Model Model


Element Z M Potential Potential  M / Z IP IP
H 1 1.008 13.610 13.718 14.250 0
He 2 4.003 24.606 49.244 49.875 35.625

Li 3 6.941 5.394 12.480 12.469 1.781


Be 4 9.012 9.326 21.011 23.327 9.077
B 5 10.811 8.309 17.966 17.055 2.805
C 6 12.011 11.266 22.551 21.570 7.320
N 7 14.007 14.544 29.101 27.281 13.031
O 8 15.999 13.631 27.260 27.281 13.031
F 9 18.998 17.438 36.810 34.504 20.254
Ne 10 20.180 21.587 43.562 43.641 29.391

Na 11 22.990 5.145 10.753 10.910 3.340


Mg 12 24.305 7.656 15.506 16.565 2.315
Al 13 26.982 5.996 12.444 14.923 0.673
Si 14 28.086 8.154 16.357 18.874 4.624
P 15 30.974 10.498 21.677 23.871 9.621
S 16 32.066 10.373 20.790 23.871 9.621
Cl 17 35.453 12.977 27.063 30.192 15.942
Ar 18 39.948 15.778 35.017 38.186 23.936

Periods 1, 2 and 3.

www.intechopen.com
156 Theoretical Concepts of Quantum Mechanics

Charge Mass Ionization IP  Ionization Model Model


Element Z M Potential Potential  M / Z IP IP
K 19 39.098 4.346 8.944 9.546 4.704
Ca 20 40.078 6.120 12.265 13.057 1.193
Sc 21 44.956 6.546 14.013 13.057 1.193
Ti 22 47.867 6.826 14.851 13.638 0.612
V 23 50.942 6.743 14.934 14.244 0.006
Cr 24 51.996 6.774 14.676 14.877 0.627
Mn 25 54.938 7.438 16.345 15.539 1.289
Fe 26 55.845 7.873 16.911 15.539 1.289
Co 27 58.933 7.863 17.163 16.229 1.980
Ni 28 58.693 7.645 16.026 16.951 2.701
Cu 29 63.546 7.728 16.934 17.705 3.455
Zn 30 65.390 9.398 20.485 18.492 4.242
Ga 31 69.723 6.006 13.509 14.494 0.244
Ge 32 72.610 7.905 17.936 17.860 3.610
As 33 74.922 9.824 22.303 22.007 7.757
Se 34 78.960 9.761 22.669 22.007 7.757
Br 35 79.904 11.826 26.998 27.116 12.866
Kr 36 83.800 14.015 32.623 33.412 19.162

Period 4.

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 157

Charge Mass Ionization IP  Ionization Model Model


Element Z M Potential Potential  M / Z IP IP
Rb 37 85.468 4.180 9.657 9.546 4.704
Sr 38 87.620 5.695 13.132 13.057 1.193
Y 39 88.906 6.390 14.567 13.057 1.193
Zr 40 91.224 6.846 15.614 13.638 0.612
Nb 41 92.906 6.888 15.608 14.244 0.006
Mo 42 95.940 7.106 16.232 14.877 0.627
Tc 43 98.000 7.282 16.597 15.539 1.289
Ru 44 101.070 7.376 16.942 15.539 1.289
Rh 45 102.906 7.469 17.080 16.230 1.980
Pd 46 106.420 8.351 19.319 16.951 2.701
Ag 47 107.868 7.583 17.403 17.705 3.455
Cd 48 112.411 9.004 21.087 18.492 4.242
In 49 114.818 5.788 13.563 14.494 0.244
Sn 50 118.710 7.355 17.462 17.860 3.610
Sb 51 121.760 8.651 20.655 22.007 7.757
Te 52 127.600 9.015 22.120 22.007 7.757
I 53 126.904 10.456 25.037 27.116 12.866
Xe 54 131.290 12.137 29.508 33.412 19.162

Period 5.

www.intechopen.com
158 Theoretical Concepts of Quantum Mechanics

Charge Mass Ionization IP  Ionization Model Model


Element Z M Potential Potential  M / Z IP IP
Cs 55 132.905 3.900 9.425 9.546 4.704
Ba 56 137.327 5.218 12.796 13.057 1.192
La 57 138.906 5.581 13.600 12.393 1.857
Ce 58 140.116 5.477 13.232 12.583 1.667
Pr 59 140.908 5.425 12.957 12.776 1.474
Nd 60 144.240 5.498 13.217 12.972 1.278
Pm 61 145.000 5.550 13.192 13.171 1.079
Sm 62 150.360 5.633 13.660 13.374 0.876
Eu 63 151.964 5.674 13.687 13.579 0.671
Gd 64 157.250 6.141 15.089 13.579 0.671
Tb 65 158.925 5.851 14.305 13.787 0.463
Dy 66 162.500 5.934 14.609 13.999 0.251
Ho 67 164.930 6.027 14.836 14.213 0.037
Er 68 167.260 6.110 15.029 14.431 0.181
Tm 69 168.934 6.183 15.137 14.653 0.403
Yb 70 170.040 6.255 15.463 14.878 0.627
Lu 71 174.967 5.436 13.395 17.860 3.610
Hf 72 178.490 7.054 17.487 18.755 4.505
Ta 73 180.948 7.894 19.568 19.696 5.446
W 74 183.840 7.988 19.844 20.684 6.434
Re 75 186.207 7.884 19.574 21.721 7.471
Os 76 190.230 8.714 21.811 21.721 7.471
Ir 77 192.217 9.129 22.788 22.811 8.560
Pt 78 195.076 9.025 22.571 23.955 9.705
Au 79 196.967 9.232 23.019 25.156 10.906
Hg 80 200.530 10.446 26.184 26.418 12.168
Tl 81 204.383 6.110 15.417 16.515 2.265
Pb 82 207.200 7.427 18.768 19.696 5.446
Bi 83 208.980 7.293 18.361 23.490 9.240
Po 84 209.000 8.423 20.958 23.490 9.240
At 85 210.000 28.015 13.765
Rn 86 222.000 10.757 27.769 33.412 19.164

Period 6.

www.intechopen.com
Better Unification for Physics in General Through Quantum Mechanics in Particular 159

Charge Mass Ionization IP  Ionization Model Model


Element Z M Potential Potential  M / Z IP IP
Fr 87 223.000 9.546 4.704
Ra 88 226.000 5.280 13.560 13.057 1.193
Ac 89 227.000 6.950 17.727 12.393 1.857
Th 90 232.038 6.089 15.699 12.583 1.667
Pa 91 231.036 5.892 14.959 12.776 1.474
U 92 238.029 6.203 16.050 12.972 1.277
Np 93 237.000 6.276 15.994 13.171 1.079
Pu 94 244.000 6.068 15.752 13.374 0.876
Am 95 243.000 5.996 15.337 13.579 0.671
Cm 96 247.000 6.027 15.507 13.579 0.671
Bk 97 247.000 6.234 15.875 13.787 0.463
Cf 98 251.000 6.307 16.154 13.999 0.251
Es 99 252.000 6.421 16.345 14.213 0.037
Fm 100 257.000 6.504 16.716 14.431 0.181
Md 101 258.000 6.587 16.827 14.653 0.403
No 102 259.000 6.660 16.911 14.877 0.627
Lf 103 17.859 3.610
Rf 104 18.755 4.505
Db 105 19.696 5.446
Sg 106 20.684 6.434
Bh 107 21.721 7.471
Hs 108 21.721 7.471
Mt 109 22.811 8.561
Uun 110 23.955 9.705
Uuu 111 25.156 10.906
Uub 112 26.418 12.168
??? 113 16.515 2.265
??? 114 17.019 2.769
??? 115 23.490 9.240
??? 116 23.490 9.240
??? 117 28.015 13.765
??? 118 33.412 19.162

Period 7.

www.intechopen.com
160 Theoretical Concepts of Quantum Mechanics

9. Acknowledgments
The author thanks colleagues for deep and intense conversations on the topics discussed
here, especially Dr. Peter Enders, Dr. Yuri Keilman, Dr. Robert Kiehn, Prof. Zbigniew
Oziewicz, and Dr. Tom Phipps.

10. References
Aspden, H. (1990) ‘Electron Clusters’ (Correspondence) Galilean Electrodynamics vol. 1, 81-82.
Beckmann, P. (1990), “Electron Clusters”, Galilean Electrodynamics vol. 1, 55-58 (1990); see also
vol. 1, 82.
Kiein, R. M. (2009), Non-Equilibrium Systems and Irreversible Processes, Vol. 4, Adventures in
Applied Topology, especially Sect. 2.2.3,
http://www22.pair.com/csdc/kok/ebookvol4.pdf.
Piestrup, M.A., Puthoff, H.E., & Ebert, P.J. (1998). Correlated Emissions of Electrons, Galilean
Electrodynamics vol. 9, 43-49.
Putz, M. V., Ed., (2009) Quantum Frontiers of Atoms and Molecules, Nova Science Publishers,
New York.
Rowlands, P. (2007), Zero to Infinity – The Foundations of Physics, Chapter 16, The Factor 2 and
Duality, World Scientific, New Jersey, London, etc.
Thomas, L.H. (1927), The Kinematics of an electron with an Axis, Phil. Mag. S. 7, 3 (13) 1-22.
Wheeler, J.A., and Feynman, R.P. (1945). Interaction with the Absorber as the Mechanism of
Radiation, Revs. Mod. Phys., 17 (2 & 3) 157-181.
Wheeler, J.A., and Feynman, R.P. (1949). Classical Electrodynamics in Terms of Direct
Interparticle Action, Revs. Mod. Phys., 21 (3) 425-433.
Whitney, C.K. (2012). Algebraic Chemistry: Applications and Origins, Nova Science Publishers.

www.intechopen.com
Theoretical Concepts of Quantum Mechanics
Edited by Prof. Mohammad Reza Pahlavani

ISBN 978-953-51-0088-1
Hard cover, 598 pages
Publisher InTech
Published online 24, February, 2012
Published in print edition February, 2012

Quantum theory as a scientific revolution profoundly influenced human thought about the universe and
governed forces of nature. Perhaps the historical development of quantum mechanics mimics the history of
human scientific struggles from their beginning. This book, which brought together an international community
of invited authors, represents a rich account of foundation, scientific history of quantum mechanics, relativistic
quantum mechanics and field theory, and different methods to solve the Schrodinger equation. We wish for
this collected volume to become an important reference for students and researchers.

How to reference
In order to correctly reference this scholarly work, feel free to copy and paste the following:

Cynthia Kolb Whitney (2012). Better Unification for Physics in General Through Quantum Mechanics in
Particular, Theoretical Concepts of Quantum Mechanics, Prof. Mohammad Reza Pahlavani (Ed.), ISBN: 978-
953-51-0088-1, InTech, Available from: http://www.intechopen.com/books/theoretical-concepts-of-quantum-
mechanics/better-unification-for-physics-in-general-through-quantum-mechanics-in-particular-

InTech Europe InTech China


University Campus STeP Ri Unit 405, Office Block, Hotel Equatorial Shanghai
Slavka Krautzeka 83/A No.65, Yan An Road (West), Shanghai, 200040, China
51000 Rijeka, Croatia
Phone: +385 (51) 770 447 Phone: +86-21-62489820
Fax: +385 (51) 686 166 Fax: +86-21-62489821
www.intechopen.com

You might also like